Department of Chemistry, Stony Brook University, Stony Brook, New York , United States

Size: px
Start display at page:

Download "Department of Chemistry, Stony Brook University, Stony Brook, New York , United States"

Transcription

1 Photo-assisted Intersystem Crossing: The predominant triplet formation mechanism in some Isolated Polycyclic Aromatic Molecules excited with pulsed lasers Philip M. Johnson a, and Trevor J. Sears, Department of Chemistry, Stony Brook University, Stony Brook, New York , United States Department of Chemistry, Brookhaven National Laboratory, Upton, New York 11973, United States ABSTRACT Naphthalene, anthracene, and phenanthrene are shown to have very long-lived triplet lifetimes when the isolated molecules are excited with nanosecond pulsed lasers resonant with the lowest singlet state. For naphthalene, triplet state populations are created only during the laser pulse, excluding the possibility of normal intersystem crossing at the one photon level, and all molecules have triplet lifetimes greater than hundreds of microseconds, similar to the behavior previously reported for phenylacetylene. Although containing 7 to 12 thousand cm -1 of vibrational energy, the triplet molecules have ionization thresholds appropriate to vibrationless T 1 states. The laser power dependences (slopes of log-log power plots) of the excited singlet and triplet populations are about 0.7 for naphthalene and about 0.5 for anthracene. Kinetic modeling of the power dependences successfully reproduces the experimental results and suggests that the triplet formation mechanism involves an enhanced spin orbit coupling caused by sigma character in states at the 2-photon level. Symmetry Adapted Cluster- Configuration Interaction calculations produced excited state absorption spectra to provide guidance for estimating kinetic rates and the sigma character present in higher electronic states. It is concluded that higher excited state populations are significant when larger molecules are excited with pulsed lasers and need to be taken into account whenever discussing the molecular photodynamics. a Corresponding Author: ; Philip.johnson@stonybrook.edu; telephone (631)

2 INTRODUCTION When energy is deposited into a molecule via a photon causing a transition to an excited electronic state, initially that new energy exists mostly as electronic energy, with only small amounts of rotational and vibrational being created by the photon. If a molecule is large enough, even if the photon is not energetic enough to cause photodissociation, the electronic energy can be partially or totally transformed into vibrational energy in a process called a radiationless transition. The quantum mechanical basis for describing the details of this energy transformation has been studied extensively 1, and it has been found that a key element in determining the behavior of the particular excited states involved in the transition is the density of states of the receiving manifold at the energy of the initial state. Because the density of vibrational states at a given location in a vibrational well goes up extremely quickly with the number of atoms in the molecule, excited state behavior has been categorized as being in either the small molecule limit or the large molecule limit, along with a very interesting intermediate case, where what happens depends dramatically on the details of the excitation. Of these three cases, it has generally been considered that the large molecule case was in many respects the simplest. The density of states in the vibrational manifolds at the energies being considered is so high that the states formed essentially a continuum, and the transition rates are averaged over so many possibilities that quantum mechanical considerations were able to be disregarded. Classical kinetic schemes have been fully capable of describing the observed phenomena. When the theories of radiationless transitions were being developed, almost all of the experiments examining how molecules evolve in their excited states were done by recording the fluorescence emission from the samples. For larger molecules (10 or so atoms or more), it was found that if a radiationless transition was from the photon-coupled state into an electronic state lower by a substantial amount (the energy gap was large), the fluorescence decay was always exponential, and 2

3 could be described as being in competition with other processes that converted energy into vibrations without photon emission. In all discussions of radiationless transitions, one needs to keep in mind the environment of the molecules in question. In solids and liquids there are phonon modes which provide a vast bath that can irreversibly receive vibrational energy from the excited molecule, and external forces that can change the interstate couplings involved in the excited state evolution. In gases with higher temperatures and densities, collisions can play a role by inducing energy transfer as well as external perturbations in the couplings. In environments where the target molecules are completely isolated (very rarified gases, supersonic beams, and outer space), when a molecule is electronically excited it has no way to rid itself of energy except by radiation. Fluorescence is a major decay route, but in the laboratory, infrared emission and phosphorescence generally take too long and are not generally observed on the time scale of the experiment. Therefore, except for fluorescence, isolated molecule excited states evolve only by rearranging the energies contained in their various electronic and vibrational modes, without any loss of total energy. Usually the exciting photon deposits energy mostly into a single electronic mode, but the vastly greater number of vibrational degrees of freedom in larger molecules dictates that populations will proceed in the direction of ever greater vibrational excitation until the electrons return to their ground state. Here we only consider the isolated molecule situation. When dealing with aromatic hydrocarbons, where spin-orbit coupling is small, it is convenient to discuss the problem in terms of zero-order pure spin states and small couplings between them. Thus for a closed-shell system we have a singlet manifold and a triplet manifold. Transitions between electronic states within a spin manifold are called internal conversion (IC), whereas transitions between spin manifolds are called intersystem crossing (ISC). The rates of IC and ISC in a classical kinetic framework are determined by Franck-Condon Factors (FCF's), spin-orbit coupling (SO), and Born-Oppenheimer 3

4 operators (BO), as well as the gaps between electronic states (Δ), and the density of vibrational levels of the lower state isoenergetic with the optically pumped level of the upper state (ρ). 2,3 Following the development of radiationless transition theories, pump-probe ionization experiments in supersonic beams enabled the measurements of the time-dependent populations of not only the singlet electronic states, but also the triplets. In a collisionless supersonic beam environment, ISC proceeds from the excited singlet state to a vibrationally excited triplet isoenergetic to the singlet. Therefore the triplet has a vibrational energy equal to the electronic energy difference between S 1 and T 1, and has no means of losing energy during the time scale of the experiment (phosphorescence is too slow). For benzene 4,5 and a variety of substituted benzenes 6,7,8 and related azines 9,10, it was found that large-molecule kinetic mechanisms involving competing IC and ISC processes in isolated molecules fit the data quite well. The singlet ionization signals decay with the fluorescence lifetime, while the triplet signals build up during this period and subsequently decay exponentially (presumably by a second ISC to the singlet ground state) with lifetimes of around a microsecond. This was taken as validation of the general theoretical framework that had been developed. More recently, studies have begun to appear which may indicate that perhaps the whole story has not been told about the excited state dynamics of larger molecules. First of all, a series of papers by Kato, Baba and coworkers 11,12,13,14 described the use of ultra-high resolution CW Zeeman spectroscopy to measure the magnitude of SO coupling in the singlet excited states of various hydrocarbons. For each molecule without a heteroatom incorporated, they reported that they could not detect enough coupling to account for a finite ISC rate. Secondly, in our laboratory it was found that phenylacetylene and benzonitrile pumped by nanosecond lasers did not follow a classical kinetic scheme at all. 15,16 Their triplet states were entirely populated during the laser pulse--only--and then the triplet populations did not decay at all during the 140 μs we could observe them in the beam. This is completely inexplicable in the generally accepted model of the way larger isolated molecules are supposed to behave. On a 4

5 different time scale, Fielding and coworkers 17,18 have found using femtosecond pulses that in benzene, ISC and IC proceeds with ultrafast rates, an unexpected result given the longer-time experiments. In this paper we present new results on naphthalene (2 rings), anthracene (3 rings, D 2h ), and phenanthrene (3 rings, C 2v ) which demonstrate that unusual behavior is not restricted to benzene with triple-bonded substituents 16, but is a general property of a larger group of hydrocarbons when examined by nanosecond pulsed lasers in a supersonic beam. Experiment The apparatus used for the recent experiments has been described previously. 15,16 All were done in a pulsed, skimmed supersonic beam of helium expanding from about 3 atm, seeded with the molecule of interest. The pulsed valve and sample chamber immediately behind it were heated to temperatures between 60 C and 150 C to provide adequate vapor pressures of the larger molecules studied here. Temperatures were typically 60 C for naphthalene, 150 C for anthracene, and 130 C for phenanthrene. The distance between the valve and the ion/electron collection point is 25 cm. Molecules seeded in the helium carrier gas were excited to their lowest singlet electronic state (usually the origin, but for naphthalene the 8 1 vibration at cm -1 was more often used because it is an order of magnitude more intense) by a YAG-pumped, doubled dye laser (about 8 ns pulse width), and subjected to a second pulse from a higher photon energy laser intended to probe the excited state population by ionizing only the excited molecules. The probe laser was either an argon fluoride excimer laser, or a YAG-pumped doubled or tripled dye laser, depending upon the wavelength desired. The two laser beams used were carefully combined to be coincident with dichroic mirrors, and counterpropagated together, unfocused, down the length of the supersonic beam. The timing between the lasers was controlled by programmable delay generators, with a timing jitter between 15 and 20 ns for the excimer/dye laser combination and <10 ns when two dye lasers were used. The timing between the pulsed valve and the probe laser was kept constant and adjusted so the leading edge of the molecular 5

6 pulse was always being sampled, avoiding excess clustering. Mechanical shutters in both laser beams allow for background subtraction during the course of data acquisition. The results reported here depend upon two types of experiments--pump-probe time delay studies with ion detection, and pump-probe time-of-flight (TOF) photoelectron spectroscopy (PES) at various time delays. The basic experiment for examining the evolution of excited state populations is a time-delay scan. Ions created by the probe laser beam were sent into a TOF mass spectrometer, and only the parent mass peak was sampled. The time between the laser firings was varied from having the probe beam fired before the pump, to having the probe follow the pump by as much as 150 μs. Data is accumulated at each delay value for typically 2 s while cycling through the set of delays many times. This reduces the effects of long-time drift in the properties of the lasers. Laser intensities were set to minimize the amount of single-laser signal due to multiphoton ionization (which is fortunately minimal for these molecules because the S 1 energy is less than halfway to the ionization energy), but the shutters were used to subtract any residual direct ionization signal. 193 nm TOF photoelectron spectra of the S 1 states of naphthalene and anthracene were recorded using a 30 cm flight tube, doubly shielded with mu-metal for magnetic isolation. No repeller voltage was used, and the electron flight times were recorded using a National Instruments transient digitizer card having a time resolution of 10 ns. Again, mechanical shutters were used for single-laser signal and background subtraction. As in time-delay scans, PES signals with various time delays were collected for short time periods and the delays were cycled through for long periods of time (up to two hours) so the spectra with different delays can be compared without being compromised by long term laser drift. The TOF data was converted to an energy scale using an E -3/2 factor (the Jacobian of the change from a time axis to an energy axis) to scale the intensities so that areas in the TOF and energy spectra are consistent. 6

7 Intensity dependences of the singlet and triplet signals were obtained by varying the pump laser intensity while recording a time-delay scan at each setting. At higher intensities, the power was changed by using a variable wave plate and a polarizer, directly measuring the power at each point with a photodiode. For lower intensities, the power was further attenuated using a set of calibrated neutral density filters. RESULTS AND DISCUSSION A representative time delay scan is shown in Figure 1. The signal shape looked the same for all wavelengths and molecules, consisting of a peak when the pump and probe lasers were overlapped in time, followed by an exponential decay to a constant value that persisted without attenuation as long as the flight time in our apparatus allowed--approximately 140 μs. The mathematical function for the signal (ignoring the influence of laser pulse shape) is therefore C = S 1 e (t t 0)/τ + T ; t t 0 (1) where t 0 is when the lasers are overlapped in time, and τ is the lifetime (determined by fluorescence) of the pumped singlet level. As will be explained in more detail later, S 1 is proportional to the population of the pumped singlet immediately following the pump pulse, and T is proportional to the population of a long-lived triplet state formed during the pump laser pulse. The values of S 1 and T were determined from each scan by convoluting the above function with a Gaussian representing the time widths of the lasers and any time jitter, then fitting the data to that function using a non-linear least-squares routine. It should be noted that there is no provision in this equation for any triplet population to be created after the laser pulse, unlike the case for benzene and various substituted benzenes studied earlier. An ionization threshold curve for the triplet component T can give considerable information about the character of the state. In order to produce the range of ultraviolet wavelengths for the probe laser necessary for this task, a variety of lasers, dyes and harmonic schemes had to be used. Since it is 7

8 not possible to maintain consistent control of the intensities and beam qualities over this variety of methods, it was decided to use the ratio T/(T+S 1 ) as the measure of the T ionization probability. This ratio is much less sensitive to intensities and beam quality than a measure of T alone, and as shown later in kinetic analyses, it is almost insensitive to laser intensity. It has the disadvantage that it depends upon both the singlet and triplet ionization cross-sections as a function of the probe wavelength, but the probe photon energy at the triplet ionization threshold excites a singlet, S 1, molecule to considerably higher than its threshold, so in general it can be expected the S 1 ionization curve will be continuous and fairly flat over a small range. Some checks on this determined it to be true in the regions of interest. Therefore ionization ratios were measured as a function of wavelength by doing time delay scans approximately every nanometer of the probe laser from the triplet ionization threshold to 198nm using doubling and tripling of dye lasers in BBO, and at 193nm using an ArF excimer laser. These ratios were repeatable between different dye lasers, different harmonic schemes, and from day to day. Time delay scans for all the molecules in this study were done using the 193nm laser for the probe photons. The triplet ionization ratios at this wavelength are given in table 1 for a group of aromatic molecules (for benzene, the only molecule on the list that follows a classical kinetic scheme for its radiationless transitions, the ratio given is Q isc /(Q isc + Q fl ), where the Q's are the quantum yields for intersystem crossing and fluorescence5). 193nm light is far enough above the ionization potentials of the triplet states of all these molecules that the ionization probability of both singlets and triplets should be substantial. The ionization ratios vary from zero to almost 0.9, thus showing that the amount of triplet formed depends in some intricate way on the excited state structure of the individual molecule. Figures 2 and 3 show the ionization ratio curves with respect to wavelength for naphthalene and anthracene, with the ionization threshold of the vibrationless T 1 state indicated (as determined by known state energies). Although it is expected that the ionization threshold of a vibrationally excited 8

9 state would be quite blue shifted due to Franck-Condon factors, it is seen that for both of these molecules the ionization threshold is very sharp--much more abrupt than one would expect for a state with cm -1 or cm -1 of vibrational energy. It is also notable that the first major threshold step is exactly at the D 0 - T 1 energy difference, and for anthracene there is almost no signal below that value, confirming that the long-lived state observed is T 1 in these molecules. There is a large and unexpected dip in the ionization curve of naphthalene centered at about 210nm (47600 cm -1 ). Interestingly, a very strong valence absorption feature appears in the calculated triplet-triplet absorption spectrum at that wavelength (see Figure 5 below). A possibility is that this valence absorption, along with internal conversion, is competing with the ionization continuum at those wavelengths and suppressing the ionization signal. Figure 4 shows the photoelectron spectra of the S 1 ionization process at various delays between the pump and probe lasers for naphthalene. The higher energy signal resulting from S 1 ionization decays with the fluorescence lifetimes (300ns), while the lower energy electrons from triplet ionization are there from the beginning and do not decay with time. Like phenylacetylene, the naphthalene triplet signal does not increase during the singlet lifetime, indicating there is no ISC in the absence of the laser field. There is more overlap between the singlet and triplet photoelectron spectra than there is in phenylacetylene 15 so the fact that all triplet signal is created during the laser pulse is less obvious. However, for all electron energies the signal can be fit to equation 1, showing there is no intersystem crossing from the post-laser singlet population. For anthracene the fluorescence lifetime is comparable to the width of the laser pulse. That, and time jitter effects, prevent an estimate of the magnitude of radiationless ISC in that molecule, as does the small amount of S 1 produced in phenanthrene. The results presented here parallel the behavior seen previously in phenylacetylene 15,16, showing that molecule is not an exception, but part of a substantial group of larger molecules exhibiting 9

10 similar excited state dynamics. The three major observations that are difficult to rationalize using current ideas about radiationless transitions in larger molecules are: 1) Triplet formation only takes place during the light pulse (at least for naphthalene). 2) The triplet population does not decay significantly over hundreds of microseconds. 3) Ionization thresholds for the triplet population are very sharp, and exactly at the ionization energy of a vibrationless level of T 1, even though the triplet molecules must have large amounts of vibrational energy. Triplet Formation In a normal molecule such as benzene (if that is indeed normal), triplet population is generated by intersystem crossing (ISC) from a singlet population and continues as long as the latter exists. The ISC rate is determined by the magnitude of spin-orbit (SO) coupling in the S 1 state, and the density of states in the triplet manifold isoenergetic to S 1. Attempts have been made to measure the magnitude of the SO coupling in the molecules studied here by using ultra-high-resolution Zeeman spectroscopy 11-14, but the results show that there is not enough mixing between the singlets and triplets in these molecules to account for any amount of ISC at all, much less the almost complete transfer seen here in phenanthrene (Table 1). In previous publications 15,16 we have considered a variety of mechanisms for triplet production (such as photon emission of various sorts), but have failed to find any evidence for their existence. Since we now know the behavior seen in PA is general, here we would like to propose a mechanism that has the possibility of explaining all the current observations--light-induced SO coupling, in which the singlet-triplet coupling takes place at the 2-photon level but population is driven down to the one-photon triplet level by stimulated emission. 10

11 An interesting aspect of all experiments done with pulsed lasers in a given experimental apparatus is that they are all done with about the same rate of population of the initial state. This is because the laser is turned up until one sees a signal, and then one stops increasing the power because multiphoton processes and detector saturation can compromise the measurement value. When the absorption cross-section of the initial transition is very small, a substantial amount of laser power must frequently be used. Given that significant laser flux is used in the experiments, one must consider whether multiple photon processes are important in the photodynamics. In the simplest cases, the curves of a multiple photon process can be identified and described by the relationship C=σI n, where C is some concentration produced by light with an intensity I, absorbed with a cross-section σ. For a multiphoton transition with no resonance intermediates, n is the number of photons involved in the transition, and is the slope of a plot of log C versus log I. However, even for transitions without intermediate resonances, this simple picture ignores depletion of the initial state and geometric effects related to saturation, so it is not often seen experimentally 19. For the experiments discussed here, the kinetics of excitation are better described by sequential one-photon steps, in which case a simple exponential relationship between the overall transition probability and intensity is valid only as a limiting case. Also, since stimulated emission must also be considered, any fully resonant multiple photon process without dissipative branches is a classical example of coupled equilibria. Once equilibrium is established, a steady state is achieved, and the intensity dependence can approach zero, indicative of saturation. If the laser pulse is weak enough or short enough that equilibrium is not achieved, or if dissipative processes occur, the intensity dependence can be anything from zero to the number of photons involved. However in general, a loglog power dependence plot has significant curvature, so an intensity dependence can only be defined over a small intensity region. A case that becomes fairly linear over large intensity spans for a two-step resonant process is one with a small initial cross-section and a much larger intermediate-to-final-state 11

12 cross-section. If there is a substantial loss mechanism at the 2-photon level the intensity dependence of the intermediate state population can become one or less, since the initial transition would be the ratelimiting step, even though it is a 2-photon process. SO coupling in aromatics has been studied previously in the context of accounting for the long phosphorescence lifetimes in condensed phases. Marian 20 has recently reviewed the state of theoretical calculations of SO for large molecules and its role in ISC. Before modern electronic structure calculations were available, Henry and Siebrand 21 made the argument that the primary sources of the magnitude of the T 1 -S 0 transition moment were π-σ* and σ-π* states coupled into T 1 by SO perturbation. More recently, Agrens et al 22 confirmed this using the multiconfigurational quadratic response method. It has also been a common theme that molecular distortions from vibrational motion are a potent source of SO coupling and its effect on ISC. 23 Thus it is generally accepted that, although π-π* states of hydrocarbons without heteroatoms usually have very small SO coupling, π-σ*, σ-π* and σ-σ* states have substantially more. These latter states are expected to be present at the two photon energy level, so mechanisms involving higher excited states must be considered. The discussion immediately above is particularly relevant here because a common characteristic of all the molecules considered here is that the S 1 - S 0 transitions have very low oscillator strengths, almost negligible in comparison to those of higher electronic states. Once an S 1 population is created, one must then consider the possibility of transitions to states one photon higher in energy than S 1, and these transitions have oscillator strengths that may be orders of magnitude larger than the initial one. If that is the case, the molecules will rapidly absorb another photon, resulting in a mixture of states at the two photon level. These will be strongly coupled to each other and to isoenergetic states in the triplet manifold because of the strong SO coupling at that level. Once populated, those mixed states can 12

13 undergo stimulated emission down to the lower triplet levels, in effect creating the equivalent of an ISC process. Whether that up-down process through the two-photon level is viable for a given molecule will depend upon the cross-sections for the S 1 absorption and the T n stimulated emission probability (which is equal to the absorption cross-section of the lower triplet at the appropriate geometry). This information is not available experimentally, but can be obtained from electronic structure calculations. Calculated Excited State Absorption Spectra To examine the proposition that optical coupling between excited states is a viable mechanism for inducing intersystem crossing, we have calculated the excited state compositions and oscillator strengths for states that are optically coupled in one photon to the S 0, S 1, T 1 and T 2 states of phenylacetylene, benzonitrile, naphthalene, anthracene and phenanthrene. For this we used the SAC-CI package in Gaussian (equivalent to the Equations of Motion (EOM) coupled-cluster method found in other electronic structure calculation collections) with an augmented, correlation-corrected, double zeta basis set (aug-cc-pvdz), at level one, and with 25 excited states of each allowed symmetry. As many as 12 σ orbitals are found within the energy range of the π orbitals of naphthalene, for example, so σ involvement in transitions of modest energy are to be anticipated. To provide a graphical representation of these results, simulated S 1 and T n absorption spectra are given in Figure 5. There the oscillator strengths are convoluted with a Gaussian band shape of 800 cm -1 width (chosen because of the experimental width of the S 3 -S 0 band in the naphthalene absorption spectrum of George and Morris 24, but much smaller than the probable width of any valence state at the 60,000 cm -1 level) and summed to give a spectrum scaled as a molar extinction coefficient ε that can be compared to experiment. In D 2h, symmetry labels change with the choice of axis system and here they follow the Mulliken convention, where the long axis of the molecule is y, the in-plane short axis is z, and out-of-plane is x. While these simulations are not expected to be identical to an experimental 13

14 absorption spectra (high Rydberg and continuum orbitals are not included in the basis set, and the degrees of lifetime broadening of higher states are not known), they provide a basis for discussing a proposed mechanism. Some experimental ultraviolet and vacuum ultraviolet spectra 24,25,26 are available to assess the quality of the calculated spectra from the ground state of each molecule, and also some triplet-triplet absorption spectra, over small energy ranges. The latter were obtained from molecules contained in low-temperature solid solutions. The vacuum UV spectrum of each molecule consists of a series of broad, intense, valence absorptions superimposed by sharper and weaker Rydberg lines. The broad structure (as well as the near-uv spectrum) is remarkably well reproduced by the SAC-CI calculations, with relative intensities close to experiment. The calculated S 0 VUV spectrum of naphthalene is shown in Figure 6, compared with an experimental spectrum adapted from Koch, Otto and Radler 25. Intensities are scaled to make the highest peaks equal since the experimental spectrum does not have a quantitative amplitude scale. It is seen that the major peak is coincident in energy, while the higher calculated energies seem to be low by about 0.2 ev. For anthracene, all apparent peaks are reproduced except for one at 7.0 ev, which is calculated at 7.3 ev (and relative intensities are still good). For phenanthrene, all experimental valence peaks are reproduced within 0.1 ev by the calculations. The triplet-triplet experimental spectra (available for naphthalene 27,28, anthracene 27,28,29, and phenanthrene 28 ) usually consist of a series of lines which we take to be a vibrational progression of around 1450 cm -1, of a single state, at an energy below the S 1 transition energy. One can estimate a solvent shift by comparing a solution UV spectrum with the corresponding vapor spectrum of each molecule. For naphthalene and phenanthrene the shifted calculated spectrum has a strong peak within 500 cm -1 of the experimental spectrum. For anthracene there is a 2500 cm -1 discrepancy. 14

15 In addition to transition rates, the SAC-CI calculations provide a quantitative measure of configuration mixing, giving an indication of the amount of σ-π mixing at the two photon level. Examination of the orbital content of each state of naphthalene showed that the 1 B 2g states (out-ofplane polarized from the one-photon levels) are all π-σ* and σ-π* as expected. However they have very little oscillator strength. The strongest transitions from S 1 are to A g (long axis) and B 1g (short axis) states. On examining orbital content, it is found that there is a substantial contribution from σ-σ* excitations for some of the states of these symmetries. Indeed, in the 30,000 to 40,000 cm -1 range of transition energies from S 1, the states with high oscillator strength have σ-σ* character in the range of 10-14%. A strong SO coupling is therefore likely, based on the conclusions of Henry and Siebrand 21. Optical transitions to and from the higher excited states of large molecules For transitions to the S 1 states of large molecules, both the initial and final states can be considered to be basically stationary, and described well by the usual Born-Oppenheimer (BO) wavefunctions. On progressing to states that have twice the S 1 energy, the density of both electronic and vibrational states at that level makes the situation much more complicated. It is important to realize that the density of electronic states at the 2-photon level is very large for these molecules. For example, for phenanthrene there are 56 total calculated singlet and triplet states between 7 and 8 ev (or 110 if you count the triplet spin sublevels separately). That means the average electronic state spacing is less than the zero point energy of most vibrations, and one could consider this energy range to be a continuous distribution of conical intersections. A rigorous treatment of the evolution of a state at the 2-photon therefore presents a significant challenge. As illustrated by the continuous nature of the optical spectrum, each valence state of a given energy is broadened and mixed with all other states near to that energy to one extent or another by BO breakdown and other perturbations such as SO coupling. However, the upper state of any optical transition to the two-photon level will be a non- 15

16 stationary superposition state consisting only of optically allowed components that will begin to evolve on the timescale of both electronic and nuclear motions. This non-stationary state can be written as a sum of BO states: Ψ f = k,j c kj (t) ψ k χ j (R k 0, Q k j )e iω kjt (2) ω kj = ω k + ω j, where ω k is the electronic frequency and ω j is the vibrational frequency; ψ k is the electronic wavefunction; χ j is the j th vibrational wavefunction of state k, with an equilibrium nuclear geometry of R k 0, and normal coordinates {Q k j }; and when the molecule is in a radiation field, to first order: c kj (t) = e t ψ m k A ψ i χ j χ i e iω kit 0 dt (3) ψ i and χ i are the electronic and vibrational wavefunctions of the initial state, and ω ki = ω kj - ω i, while A is the vector potential of the radiation. It is seen that having a finite c kj (t) depends upon having a finite oscillator strength and Franck- Condon factor, so the radiation field prepares an initial higher state only using components of the final eigenstate with these characteristics. Following the initial preparation, the superposition state will dephase and evolve at the whims of small intramolecular perturbations and the changed interaction potential between the electrons and nuclei. Due to the localized nature of the initial vibrational wavefunction, the initial superposition state will have the same geometry as the lower state, but the equilibrium geometry of Ψ f will not be the same as Ψ 0 and the molecule can travel away from that location in nuclear space on a short timescale, mixing in other electronic states as it goes. When each level is resonant in a multiple photon process, the steps can be considered as a sequential linear rate scheme. However, unless there is a very rapid dissipative process in some upper level, such as direct dissociation, one must also consider stimulated emission downward from each 16

17 photon level, since the rate of a backward transition is equal to the forward rate. Particularly when an upper state is highly mixed, all lower states are in play for a downward destination that meet the energy requirements, not just the states visited on the way up. Therefore, components of the superposition state that did not carry oscillator strength from the set of upward transitions can initiate transitions downward to states at fewer-photon levels that cannot be populated directly from the ground state. To make the process more definite for the present purposes, it is convenient to introduce a classical rate scheme that can be used to examine the consequences of the various transition rates in the molecules considered here. A plausible scheme of minimal complexity is shown in Figure 7. Following a transition to S 1 with a cross-section σ 1,the molecule rapidly establishes another equilibrium with a nonstationary singlet state at the two-photon level with cross-section σ 2, as well as having the possibility of undergoing ISC to the triplet state at the one photon level with a rate k 2 and fluorescing with a rate k 1. Meanwhile, some components of the singlet two-photon superposition state can establish a different equilibrium (ISC) to the triplet manifold with a rate k 3. There are dissipation channels to a sink in the singlet manifold (k 5 ) that allows for losses by internal conversion at the 2- photon level, and also by ionization (with a cross-section σ 3 ). In the triplet manifold the one- and 2- photon levels are connected by a cross-section σ 4, and an initial state populated at the one-photon level is connected to a relaxed state by internal relaxation or IC (k 4 ). All cross-sections can be considered to be first order in the laser intensity, and therefore the rate is represented by the product of a crosssection and a time-dependent intensity, k n =σ n I(t). Here we will assume the laser pulse has a Gaussian intensity dependence centered on a time t 0, giving I(t) = e (t t 0) 2 /γ 2, where γ=0.926 W, W being the full width at half maximum of the pulse (a trial inclusion of a rapid oscillation superimposed on the Gaussian to simulate mode mixing in the unseeded YAG laser did not alter any results). This scheme can be represented by a set of coupled rate equations: 17

18 da dt = e (t t 0) 2 /γ 2 σ 1 [A B] + k 1 B (4) db dt = e (t t 0) 2 /γ 2 [ σ 1 (A B) σ 2 (B C)] (k 1 + k 2 )B (5) dc dt = e (t t 0) 2 /γ 2 [σ 2 (B C) σ 3 C] (k 3 + k 5 )C + k 3 D (6) dd dt = e (t t 0) 2 /γ 2 [σ 4 (E D)] + k 3 C k 3 D (7) de dt = e (t t 0) 2 /γ 2 [σ 4 (C E)] k 4 E + k 2 B (8) dg dt = k 4E (9) dz dt = σ 5e (t t 0) 2 /γ 2 C + k 5 C (10) where A= [S 0 ], B=[S 1 ], C=a singlet superposition state at the 2 photon level, D= a mixed triplet state at the 2 photon level, and E= a triplet state at the one-photon level (T 1 or T 2 ). Z (not shown in Figure 7) is a sink representative of products of IC and ionization from the 2-photon level. It was also found necessary to include another sink at the one-photon triplet level, state G, that represents irreversible relaxation there. G is the long-lived triplet state seen in the experiment. This equation set was numerically integrated (using the Runge-Kutta method) as a function of time, assuming all population is initially in A, and the changing relative populations determined for all levels as a function of laser intensity. Initial estimations for the cross-sections were obtained from the SAC-CI calculations, and the magnitudes and trends in the population ratios compared to experimental results. One way of looking at the kinetic scheme described above is that, for the most part, it is a systematic method of introducing additional basis functions into the description of the electronic structure of a molecule being significantly polarized by a radiation field. Electrons respond much more rapidly to the light than the nuclei and the electronic states are driven into equilibrium before the 18

19 nuclei. The various equilibria and relaxation processes at different levels of excitation then distribute the population into various electronic states at the end of the pulse, at which time internal interactions take over directing the state evolution. This makes the distribution of products intensity dependent in a very complicated way that is not directly correlated to the number of photons involved in the process. In this picture levels C and D are the pure singlet or triplet nonstationary components of a mixed dressed state with both singlet and triplet character, and the 2-photon ISC rate k 3 is the dephasing rate of the components. States C, D, and E can all be considered as components of the evolving, radiationpolarized wavefunction of the molecule during the course of the pulse, as represented by equation 2. This description of the dynamics at the 2-photon level is undoubtedly simplistic, and is the weakest part of this kinetic model. Transitions downward from the two-photon level into the lower triplet manifold can terminate in any electronic state whose potential minimum is below the one-photon energy. The SAC-CI calculations indicate that there are either two or three of those states at the ground state geometry for the molecules studied here. States just above the one photon energy at the S 0 geometry could well drop below at a different geometry or when a zero-point correction is applied, so they could be considered to be lower state candidates also. For all of the molecules but anthracene, T 1 has a different symmetry than S 1, and there is a triplet state of S 1 symmetry just below the S 1 energy that can be reached from the 2-photon superposition state given reasonable Franck-Condon overlap. The energy gap between the lower triplet state energy and the S 1 energy must eventually appear in any T 1 product as vibrational energy for conservation purposes. If the downward transition happens before any IC can take place, the Franck-Condon factors (FCf's) for the transitions would favor lower triplet electronic states near to the S 1 energy. Alternately, considerable electronic energy can be transferred in the upper level to vibrations. In that case, since there are a very large number of states in 19

20 these molecules just below the two photon level, given time, the molecule could start to cascade down through these electronic states, transferring energy to the nuclear motion while keeping the total energy in the molecule the same. This relaxation would inhibit the maintenance of an equilibrium with the singlet levels however, and is explicitly included as IC in the kinetic scheme listed above, only as a sink for population. The above kinetic scheme can be used to predict the S 1 and T 1 populations at the end of the laser pulse, as a function of the intensity of the light, and these population-versus-intensity curves can be compared to measured signals. Estimated rates from absorption spectra can serve as initial values for some cross-sections, and some rates are constrained by experimental observations. For example, the fact that no triplet is formed in naphthalene after the laser pulse, forces k 2 to be zero. Also, it can be anticipated that the 2- photon IC, k 5, will be relatively large from general experience in resonance enhanced multiphoton ionization spectroscopy (REMPI) 30,31. In that method, for larger molecules the only states seen as twophoton resonances are Rydberg states. 32 This is because IC competes with ionization for valence resonances in the upper singlet manifold even though quite strong laser intensities are used in REMPI. For naphthalene, there does not seem to be any detectable REMPI spectrum at all in the S 1 region below the 2-photon IP 33, and there was negligible ionization seen from the pump laser in the present experiments, indicating k 5 is substantial. For naphthalene, the absorption calculations indicate that the ratio of oscillator strengths at the one- and two-photon levels is about 200. This, however, neglects vibronic effects in S 1 and overlapping bands at the 2-photon level. Also the 8 1 vibrational band used for most measurements is a factor of 10 more intense than the origin, so a σ 2 /σ 1 ratio of is a reasonable initial estimate. σ 4 is expected to be about double σ 2. Starting from those estimates, the fitted S 1 and T 1 intensity curves for 20

21 naphthalene are shown in Figure 8. The rate equation values used to reproduce the experimental curves are: σ 1 =4x10 7 ; σ 2 =2.5x10 9 ; σ 3 =0 ; σ 4 = 5x10 9 ; k 2 =0; k 3 =2x10 7 ; k 4 =2x10 8 ; and k 5 =2.9x10 8. Since the absolute values of the laser intensities are not known, their units in the program are scaled such that the values are exactly one at the maximum intensity, and the rates resulting from them have the values of the σ's at that point. This scheme has the additional advantage that all kinetic parameters and crosssections have units of s -1 since the intensities are represented by dimensionless ratios. The above values are all quite reasonable numbers, but should not be considered definitive since they are quite dependent on the kinetic model, and may not even be a unique solution to the one used. The results show, however, that this simple kinetic scheme is capable of reproducing the experimental trends. Other, simpler schemes that were tried could not simultaneously reproduce the curvatures, the slopes and the magnitudes. Slopes on a log-log plot of the experimental data for naphthalene versus intensity gave power dependences of 0.67 for the singlet and 0.71 for the singlet, again emphasizing the important role of the IC at the two-photon level and the difference in the singlet cross-sections. The most challenging feature in the intensity dependence curve to reproduce in the kinetics is the positive curvature of the triplet curve at low intensities. Since in a kinetic scheme the states are sequentially populated, if the triplet population has to go through the 2-photon level (versus crossing directly at the one-photon level) there is naturally a negative curvature of the triplet signal with intensity because of a quadratic intensity component. ISC at the one photon level is precluded for naphthalene because experimentally it is found there is no triplet formation after the laser pulse, which would happen if k 2 had any significant magnitude, given the long fluorescence lifetime. Together, the large σ 2 /σ 1 ratio and IC in the 2-photon singlet redistribute the populations in such a way that a positive curvature is obtained. 21

22 For anthracene, even though the S 0 - S 1 oscillator strength is much larger than in naphthalene, since a strong vibronic band was used in the latter, the S 0 - S 1 cross-sections are comparable. The intensity curves are also quite similar, with an initial rise followed by a fairly linear increase for the triplet, and a somewhat curved increase for the singlet. Because of a very low power dependence in this molecule, it is possible to attenuate the pump laser considerably and still see signal, providing measurements over three orders of magnitude of intensity and making a log-log plot desirable, as in Figure 9. There the singlet and triplet data points are accompanied by a linear least-squares fit to each logarithmic data set (black lines), and a fit from the kinetic equation calculations (dashed lines). The slopes of the least-squares lines are 0.47 for the singlet and 0.52 for the triplet. The SAC calculated absorption spectra only show a modest singlet cross-section to the twophoton level, and indicate σ 1 σ 2. Due to large uncertainties in the position of a single, large, calculated triplet-triplet absorption peak shown in Figure 5, it is difficult to estimate σ 4 from the simulated spectra. Also, for anthracene, the fluorescence lifetime is only a small amount larger (~18ns) 34 than the laser pulse width, so it is not possible to estimate the value of k 2 from whether triplet is formed after the laser pulse. The additional flexibility in allowing a finite value of k 2 is useful, and a reasonable fit to the experimental data over the entire intensity range is obtained with the kinetic values: σ 1 =2 x10 8 ; σ 2 =2x10 10 ; σ 3 =0 ; σ 4 = 2x10 7 ; k 2 =6x10 6 ; k 3 =4x10 8 ; k 4 =1x10 9 ; and k 5 =1x10 9. With the exception of k 2 (S 1 ISC), all the k n are similar to naphthalene. The very small amount of one-photon ISC (only about 1/10th of the fluorescence rate) is needed to provide enough triplet population at low intensities. At higher intensities the triplet population is provided by the 2-photon ISC (k 3 ), which is much larger than k 2, and the upper parts of the curves can be reproduced well without using k 2. The fraction of population going into IC from the 2-photon level varies considerably over the intensity range, going from 0.01 at the lowest power to 0.95 at the highest. It is primarily this sink that creates the very low power dependence exponent. Also, it is seen that the σ 1 and σ 2 values emerging from the fit are not what one would expect 22

23 from the calculated absorption spectra (i.e. not about equal). This may be due to deficiencies in the electronic structure calculations, in the kinetic scheme, or both, and needs to be examined further. However, since the experimental log-log plot is surprisingly linear over such a large intensity range, indicative of a rate-limiting step, an error in the SAC-calculated σ 2 /σ 1 ratio is suspected. This is most likely due to the fact that calculations were only done in D 2h symmetry, which eliminates S 1 -S 2 vibronic mixing that would introduce the same strong singlet transitions seen in naphthalene. Since a singlegeometry calculation uses up to two months of cpu time, a full vibronic calculation is not practical at the present time. The number of adjustable parameters in the kinetic scheme is fairly large and it would seem that almost any data set should be able to be reproduced. This, however, is not the case because the ranges are highly constrained by experimental and theoretical considerations and, particularly for anthracene, the large number of experimental points over a large intensity range. It is quite challenging to find a set of parameters that achieve a satisfying reproduction of the experimental results over the entire intensity range, but many parameters are not completely linearly independent. Thus the values given should not be construed as proven rate measurements, but as a demonstration that a kinetic model can be viable. A comparison between phenylacetylene and benzonitrile further illustrates how the calculated absorption spectra can help to explain experimental results. The former has a substantial triplet ratio (with 193 nm ionization) of 0.25, while the isoelectronic benzonitrile's ratio is only Looking at the absorption curves in Figure 5, it is seen that at the energy of the light used to pump the origin of S 1 is below all the strong absorptions of S 1 benzonitrile to the 2-photon level, while for phenylacetylene, the light energy is in the middle of several strong bands. There is also a noticeable similarity between the S 1 and the T 2 absorption curves, not surprising since those two states have almost identical sets of 23

24 configurations. Therefore whenever a superposition state gets formed at the 2-photon level it could immediately be stimulated down to the T 2 level without even changing its electronic structure. Only the spin needs to flip. The power dependence of the triplet signal of phenylacetylene was previously measured to be 1.12, 15 so in this small sample of molecules the power dependence scales inversely with the size of the molecule (and thus the density of electronic states at the 2-photon level). A more sophisticated way to model the evolution of excited electronic states is the multistate density matrix approach 35. This method can correctly take into account the coherences possible in a laser driven process, as is often necessary for atoms and small molecules. Most studies have indicated that for large molecules and complex kinetic processes, inherent relaxation and interference effects destroy any coherence before it can build up 36,37, so results are the same as when using kinetic rate equations. However, for completeness, the above kinetic scheme was implemented using density matrices, with appropriate estimates of various relaxation rates. In agreement with previous indications, there was very little difference in the cross-sections and rates obtained in fitting the experimental data using density matrices versus using rate equations. Triplet State Properties For the molecules studied here some of the molecules end up in a highly vibrationally excited triplet state when the laser pulse ends. The vibrational energy in the triplet is not a thermal distribution, but a finite number of vibrational levels that are isoenergetic with the pumped S 1 level within the transform bandwidth of the molecular rovibronic state that is excited in a particular molecule. Using the method of Haarhoff 38 (his equation 2), the experimental energy gaps, and calculated vibrational frequencies it is possible to calculate the density of triplet states per cm -1 at the S 1 level, as given in table 1. Using the experimental singlet lifetimes to calculate the transform bandwidth, we can determine the approximate number of triplet states that can couple to a single S 1 level for each molecule. For most of 24

25 these molecules, Δν Rad 1X10-4 cm -1, so the number of coupled states ranges from about 10 for benzene to almost for anthracene. In spite of considerable discussion about an "energy gap" rule 39, when comparing molecules for the purposes here, where the energy gaps are not dissimilar, the main determinant in the density of states is the number of atoms. That is because this number, in turn, determines the number of vibrational modes, which appears as the exponent of the energy gap in the Haarhoff formula. Thus phenylacetylene has about 4X10 5 coupled states after the laser pulse, explaining the difference between its behavior and that of benzene, which has only a few. During the laser pulse, one must use the transform width of the laser to determine how many states are coupled. For an 8 ns laser pulse the transform width is about 6X10-4 cm -1, so the number of coupled states will be somewhat larger than when using the radiative width, but for most of these molecules the difference is scarcely of importance. However for picosecond and femtosecond pulses the transform width encompasses many other rovibronic states and the picture changes considerably. From table 1 we see that the total vibrational energies arising from the S 1 -T 1 gap can be 7500 to cm -1, which means that the vibrational energy contents of triplet states isoenergetic to S 1 are thousands of wavenumbers past the point where vibrational normal modes have any relevance. Beyond a few thousand wavenumbers of excitation the anharmonic coupling between modes makes it is necessary to go over to a local mode picture of the vibrational motions. 40 At high vibrational excitation, one must consider many different local modes excited, all with different phases, so the atoms are moving around in almost a random fashion with respect to one another. And at the energies considered for the present molecules, there will be 10 5 to of those modes, all within an energy range of a few hundredths of a wavenumber. When considering the wavefunction of a molecule it is common to make the Born-Oppenheimer (BO) approximation, where the electronic and vibrational parts of the wavefunction can be considered 25

26 to be separable. The physical justification of the BO approximation is that the electrons, being much lighter than the nuclei, can adiabatically follow the motion of the nuclei, and the nuclear coordinates can be considered as a fixed parameter in the electronic wave functions at a given nuclear geometry (the adiabatic approximation). There are situations where the BO approximation is not valid, such as for regions of nuclear coordinate space where electronic potential energy surfaces are close to one another. In most systems, terms of the Hamiltonian ignored in the BO approximation are those responsible for the transfer of energy between electronic and vibrational motion described by BO wavefunctions. In other situations, this adiabatic approximation becomes irrelevant. For Rydberg states with large principal quantum numbers the Rydberg electron is going slowly with respect to the nuclear motion, and is at such a large distance that the moving nuclei and core electrons are indistinguishable from a point charge. Then the Rydberg electron orbits do not depend at all on the pattern of nuclear motion, and behave entirely like atomic orbitals. We would like to suggest that another situation where the adiabatic approximation is irrelevant is in the highly excited vibrational states of large triplet molecules populated by ISC. In a superposition vibrational state with components and cm -1 of energy, the nuclei are moving rapidly and randomly with respect to one another. Any electron in a region of space sufficiently removed from any atom core that it sees the electric potential of several nuclei simultaneously would be subject to a rapidly varying and randomly changing potential. The electron would essentially decouple its motion from the nuclei, forming a wavefunction with respect to the average positions, in the same way a pendulum weakly coupled to a forest of other pendula oscillating with different phases and frequencies would move independently. The unpaired electrons for the T 1 states of the molecules discussed here are in pi orbitals that have no density in the plane of the nuclei, and have exposure to a number of nuclei at each point the wavefunction has significant amplitude. 26

27 The justification for this supposition is that it would shed light on the unexplained behaviors of the triplet states of these larger molecules--that the states have very long lifetimes and that their ionization thresholds are sharp. If the unpaired electrons in the triplet states were decoupled from the nuclei, there would not be enough interaction between them to enable ISC to the electronic ground state. Also, if the electron is mostly ignoring the nuclei then FCf's would be diagonal, and there would be an abrupt ionization threshold, as observed. CONCLUDING REMARKS Given the disparity between the S 0 -S 1 and the S 1 -S 2 absorption cross-sections in naphthalene and phenanthrene, it would be surprising if states at the 2-photon level were not involved in ISC. The conclusion that ISC at the one-photon level is either absent or small in naphthalene agrees well with the observations of Baba and coworkers that ISC is negligible. In their work CW lasers were used, so higher order processes are improbable, and at the one-photon level it is likely that indeed no triplet formation happens. The small amount of SO coupling arising from the kinetic model for anthracene may not be enough to show a significant Zeeman effect from the triplet, but that will have to be examined further. However, there is still a considerable conflict between their conclusion from Zeeman work that benzene has no S 1 singlet-triplet coupling 41, and the experimental pulsed laser results 15 that show a definite ISC after the laser pulse. One would have to conclude that there must be an additional effect that makes a CW Zeeman measurement different than a nanosecond pulsed experiment. An interesting question is how much this understanding will contribute to interpreting the behavior of large molecules when excited with shorter picosecond and femtosecond pulses. Although the total photon flux in a given ultrafast pulse is less than in a nanosecond pulse, the peak powers of the lasers used in ultrafast experiments are orders of magnitude higher. Therefore, multiphoton process would be expected to be prevalent. Studies by Weber and coworkers 42 found that in phenol and naphthalene, femtosecond pulses efficiently created states at the 2-photon level (which they called 27

28 "superexcited" states) that they were able to study by time-delayed photoelectron spectroscopy. Clear signatures of high Rydberg states were seen, but there was also a broad background that may be due to valence components of a highly mixed state. The work by Fielding and coworkers 17,18 on benzene reports an ultrafast ISC rate in the "channel three" upper vibrational region of S 1. This is in contrast to the very slow (microsecond) ISC rates seen in our previous work on lower vibrations of S 1 benzene5,15. In their experiment, the excitation energy is resonant with the upper part of the Franck-Condon envelope for the S 1 state, a region of very weak absorption. Thus it is possible that the 2nd photon absorption cross-section is considerably higher than the initial one, and the linear intensity dependence seen is due to a rate-limited kinetic process. Rapid ISC at the 2-photon level is easier to rationalize, but if the reported rapid relaxation represents the 2-photon dephasing rate, the femtosecond pulses do not last long enough to stimulate the molecule down into the lower triplet states. Thus the low energy of the electrons in the photoelectron spectra cannot be easily explained when invoking a higher energy ISC. On the other hand, if a dominant part of a two-photon state consists of two-electron-excited configurations, low energy photoelectrons could occur because the primary ionization transitions would be to electronically excited cations. It would be of interest to explore this further. Ultrafast experiments on the polycyclic aromatics could well be instrumental in further elucidating the roll of higher states in the dynamic process in molecules following pulsed laser excitation. Because of their large polarizabilities and relatively weak S 1 transitions, phenylacetylene and the polycyclic aromatics may represent an extreme in the range of molecular behaviors. However, there are a great number of molecules in this category, and for these the disregard of multiphoton processes is ill advised. Due to the complexity of the kinetic process, a simple power dependence measurement is insufficient evidence that higher states are not involved in the evolution of populations when pulsed lasers are used. We have not been concerned with the situation where very low powers are used, such as in ultrahigh resolution studies where CW lasers are necessary. However, even though short singlet 28

29 lifetimes make 2-photon excitation in the singlet unlikely at low intensity, that does not mean multiple photon effects are impossible. If there is ISC at the one-photon level, a long-lived triplet could well be pumped back into the singlet, thus decreasing the triplet population from what one would predict from a given ISC rate. Ultimately for larger aromatics, it seems likely that the comfortable picture of excited state dynamics between pure-spin BO states will have to be abandoned if the details of the process are to be understood. Only S 0 and S 1 are in that category, and better descriptions of all other states in a radiation field will possibly involve something like a density functional approach. The idea would be that there is a dressed state of the molecule characterized by a time-dependent electron density, which is connected to various irreversible pathways that lead to measurable product state populations. Hopefully this will also enhance the intuition that enables generality and insight. ACKNOWLEDGMENTS We gratefully valuable discussions with Prof. Thomas Weinacht concerning the dynamics of molecular excited states. TJS is supported at Brookhaven National Laboratory under Contract Nos. DE-AC02-98CH10886 and DE-SC with the U.S. Department of Energy and supported by its Office of Basic Energy Sciences, Division of Chemical Sciences, Geosciences, and Biosciences. 29

30 Table 1 Some properties of the molecules discussed in this work. All but benzene exhibit light-induced ISC. All energies are in wavenumbers, and the densities of states are in inverse wavenumbers. Molecule S 1 Energy T 1 Energy Ionization Energy S 1 -T 1 Gap Triplet Ratio- 193nm Probe Density Of T 1 States at S 1 S 1 Expt. Osc. Str. S 1 Theo. Osc. Str. Benzene a b c d 1.01X e 0 Phenylacetylene f g h g 3.9X i Benzonitrile 36513g j k X l Naphthalene-origin m n 65687n X o Naphthalene X10 10 Anthracene p q r X s Phenanthrene t u v X u a A. Sur, J. Knee and P. Johnson, J. Chem. Phys. 77, (1982). b S. Sharpe and P. Johnson, J. Chem. Phys. 81, (1984). c L. Chewter, K. Muller-Dethlefs and E. Schlag, Chem. Phys. Lett. 135, (1987). d From Q ISC /(Q fl +Q ISC ); C. Otis, J. Knee, and P. Johnson, J. Phys. Chem. 87, (1983). e M. Goeppert-Mayer and A. Sklar, J. Chem. Phys. 6, 645 (1938). f J. Hofstein, H. Xu, T. Sears and P. Johnson, J. Phys. Chem. A 112, (2008). g EELS solid; P. Swiderek, B. Gootz and H. Winterling, Chem. Phys. Lett. 285, 246 (1998). h C. Kwon, H. Kim and M. Kim, J. Phys. Chem. A 107,10969 (2003). i G. W. King and S. P. So, J. Mol. Struct. 37, 543 (1971). j K. Takei and Y. Kanda, Spectrochim. Acta 18, (1962). k M. Aarak, S. Sato and K. Kimura, J. Phys. Chem. 100, (1996). l T. Peacock and P. Wilkinson, Proc. Phys. Soc. 79, 105 (1962). m M. Cockett, H. Ozeki, K. Okuyama and K. Kimura, J. Chem. Phys. 98, (1993). n H. Gatterman and M. Stockburger, J. Chem. Phys. 63, (1975). o G. George and G. Morris, J. Mol. Spect. 26, (1968). p M. Baba, M. Saitoh, K. Taguma, K. Shinohara, K. Yoshido, Y. Semba, S. Kasahara, N. Nakayama, H. Goto, T. Ishimoto and U. Nagashima, J. Chem. Phys. 130, (2009). q J. Schiedt and R. Weinkauf, Chem. Phys. Lett. 266, 201 (1997). r M. Cockett and K. Kimura, J. Chem. Phys. 100, (1994). s J. Ferguson, L. Reeves and W. Schneider, Can. J. Chem. 35, (1957). t J. Warren, J. Hayes and G. Small, Chem. Phys. 102, (1986). u EELS; P. Swiderek, M. Michaud, G. Hohlneicher and L. Sanche, Chem. Phys. Lett. 176, (1991). v J. Piest, J. Oomens, J. Bakker, G. von Helden and G. Meijer, Spectrochim. Acta A 57, (2001). 30

31 Figure Ion Signal (a.u.) Laser Delay-ns Figure 1 A time-delay scan of the pump-probe ionization of naphthalene excited to the 8 1 vibration of S 1 and ionized at a later time by 193nm photons. An initial population is composed of both singlets and triplets. The singlet component decays with the fluorescence lifetime, while the triplet signal persists until the excited sample passes the detection region. There is a time scale change after the singlet decay to accommodate the much longer triplet lifetime. 31

32 Triplet Percent Figure Naphthalene Triplet Ionization Threshold 15 D 0 <-- T Wavelength of Ionization Photon--nm Figure 2: Ratio of the naphthalene triplet ionization signal to the total ionization signal when the pump and probe lasers are overlapped in time, as a function of the wavelength of the probe laser. The vibrationless T 1 ionization energy is shown. 32

33 Triplet Percent Figure Anthracene Triplet Ionization Threshold D 0 <--T Figure 3 As in figure 2, but for anthracene triplet. Wavelength of Ionization Photon--nm 33

34 Figure 4 Naphthalene S 1 8 PES--193nm Laser Delays 20ns 200ns 330ns 630ns 2000ns 5000ns Electron Energy (ev) Figure 4 The pump-probe photoelectron spectrum of naphthalene excited to the 8 1 vibration of S 1, and ionized with 193 nm photons at various different pump-probe decays. The singlet undergoes transitions to both the ground and first excited states of the ion, and the signal decays exponentially to the residual triplet signal seen at long delay time. 34

35 Figure 5 Molar Extinction Coefficient (l/mol-cm) Naphthalene S 1 Absorption T 1 T Wavenumbers Molar Extinction Coefficient (l/mol-cm) Anthracene S 1 B 1u Absorption T 1 B 1u T 2 B 3g T 3 B 2u Wavenumbers Molar Extinction Coefficient (l/mol-cm) Phenanthrene S 1 Absorption T 1 T Wavenumbers Molar Extinction Coefficient (l/mol-cm) Phenylacetylene S 1 Absorption T 1 T Wavenumbers Molar Extinction Coefficient (l/mol-cm) Benzonitrile S 1 Absorption T 1 T Wavenumbers Figure 5 Absorption spectra for the lowest singlet state and lower triplet states of naphthalene, anthracene, phenanthrene, phenylacetylene, and benzonitrile calculated using the symmetry adapted cluster-configuration interaction method. The energies of photons resonant with S 1 -S 0 are indicated. To reduce congestion, spectra are offset from one another by molar extinction coefficient units. Absorption values are calculated from state oscillator strengths using Gaussian line shapes of 800 cm -1 width. 35

36 Figure 6 Naphthalene VUV Absorbance Electron Volts Figure 6: A comparison between the experimental vacuum ultraviolet absorption spectrum of Koch, Otto, and Radler (solid line), and a spectrum derived from SAC-CI calculations (dashed line). 36

37 Figure 7 Figure 7: The multistate kinetic scheme used to model the population dynamics of isolated polycyclic aromatic molecules irradiated by a nanosecond laser pulse. Sigmas represent absorption and stimulated emission cross-sections, while k's are the rates of various intrinsic relaxation processes. State labels are characterized in the text. 37

Laser Dissociation of Protonated PAHs

Laser Dissociation of Protonated PAHs 100 Chapter 5 Laser Dissociation of Protonated PAHs 5.1 Experiments The photodissociation experiments were performed with protonated PAHs using different laser sources. The calculations from Chapter 3

More information

CHAPTER 13 Molecular Spectroscopy 2: Electronic Transitions

CHAPTER 13 Molecular Spectroscopy 2: Electronic Transitions CHAPTER 13 Molecular Spectroscopy 2: Electronic Transitions I. General Features of Electronic spectroscopy. A. Visible and ultraviolet photons excite electronic state transitions. ε photon = 120 to 1200

More information

PHOTOCHEMISTRY NOTES - 1 -

PHOTOCHEMISTRY NOTES - 1 - - 1 - PHOTOCHEMISTRY NOTES 1 st Law (Grotthus-Draper Law) Only absorbed radiation produces chemical change. Exception inelastic scattering of X- or γ-rays (electronic Raman effect). 2 nd Law (Star-Einstein

More information

What dictates the rate of radiative or nonradiative excited state decay?

What dictates the rate of radiative or nonradiative excited state decay? What dictates the rate of radiative or nonradiative excited state decay? Transitions are faster when there is minimum quantum mechanical reorganization of wavefunctions. This reorganization energy includes

More information

Figure 1 Relaxation processes within an excited state or the ground state.

Figure 1 Relaxation processes within an excited state or the ground state. Excited State Processes and Application to Lasers The technology of the laser (Light Amplified by Stimulated Emission of Radiation) was developed in the early 1960s. The technology is based on an understanding

More information

Chemistry 2. Molecular Photophysics

Chemistry 2. Molecular Photophysics Chemistry 2 Lecture 12 Molecular Photophysics Assumed knowledge Electronic states are labelled using their spin multiplicity with singlets having all electron spins paired and triplets having two unpaired

More information

LABORATORY OF ELEMENTARY BIOPHYSICS

LABORATORY OF ELEMENTARY BIOPHYSICS LABORATORY OF ELEMENTARY BIOPHYSICS Experimental exercises for III year of the First cycle studies Field: Applications of physics in biology and medicine Specialization: Molecular Biophysics Fluorescence

More information

Chapter 15 Molecular Luminescence Spectrometry

Chapter 15 Molecular Luminescence Spectrometry Chapter 15 Molecular Luminescence Spectrometry Two types of Luminescence methods are: 1) Photoluminescence, Light is directed onto a sample, where it is absorbed and imparts excess energy into the material

More information

NPTEL/IITM. Molecular Spectroscopy Lectures 1 & 2. Prof.K. Mangala Sunder Page 1 of 15. Topics. Part I : Introductory concepts Topics

NPTEL/IITM. Molecular Spectroscopy Lectures 1 & 2. Prof.K. Mangala Sunder Page 1 of 15. Topics. Part I : Introductory concepts Topics Molecular Spectroscopy Lectures 1 & 2 Part I : Introductory concepts Topics Why spectroscopy? Introduction to electromagnetic radiation Interaction of radiation with matter What are spectra? Beer-Lambert

More information

Chem 442 Review of Spectroscopy

Chem 442 Review of Spectroscopy Chem 44 Review of Spectroscopy General spectroscopy Wavelength (nm), frequency (s -1 ), wavenumber (cm -1 ) Frequency (s -1 ): n= c l Wavenumbers (cm -1 ): n =1 l Chart of photon energies and spectroscopies

More information

Abstract... I. Acknowledgements... III. Table of Content... V. List of Tables... VIII. List of Figures... IX

Abstract... I. Acknowledgements... III. Table of Content... V. List of Tables... VIII. List of Figures... IX Abstract... I Acknowledgements... III Table of Content... V List of Tables... VIII List of Figures... IX Chapter One IR-VUV Photoionization Spectroscopy 1.1 Introduction... 1 1.2 Vacuum-Ultraviolet-Ionization

More information

Excited States Calculations for Protonated PAHs

Excited States Calculations for Protonated PAHs 52 Chapter 3 Excited States Calculations for Protonated PAHs 3.1 Introduction Protonated PAHs are closed shell ions. Their electronic structure should therefore be similar to that of neutral PAHs, but

More information

Modern Optical Spectroscopy

Modern Optical Spectroscopy Modern Optical Spectroscopy With Exercises and Examples from Biophysics and Biochemistry von William W Parson 1. Auflage Springer-Verlag Berlin Heidelberg 2006 Verlag C.H. Beck im Internet: www.beck.de

More information

5.74 Introductory Quantum Mechanics II

5.74 Introductory Quantum Mechanics II MIT OpenCourseWare http://ocw.mit.edu 5.74 Introductory Quantum Mechanics II Spring 009 For information about citing these materials or our Terms of Use, visit: http://ocw.mit.edu/terms. Andrei Tokmakoff,

More information

UV-vis (Electronic) Spectra Ch.13 Atkins, Ch.19 Engel

UV-vis (Electronic) Spectra Ch.13 Atkins, Ch.19 Engel XV 74 UV-vis (Electronic) Spectra-2014 -Ch.13 Atkins, Ch.19 Engel Most broadly used analytical tech / especially bio-applic. inexpensive optics / solvent & cell usually not problem intense transitions

More information

Photoelectron Spectroscopy using High Order Harmonic Generation

Photoelectron Spectroscopy using High Order Harmonic Generation Photoelectron Spectroscopy using High Order Harmonic Generation Alana Ogata Yamanouchi Lab, University of Tokyo ABSTRACT The analysis of photochemical processes has been previously limited by the short

More information

Singlet. Fluorescence Spectroscopy * LUMO

Singlet. Fluorescence Spectroscopy * LUMO Fluorescence Spectroscopy Light can be absorbed and re-emitted by matter luminescence (photo-luminescence). There are two types of luminescence, in this discussion: fluorescence and phosphorescence. A

More information

What the Einstein Relations Tell Us

What the Einstein Relations Tell Us What the Einstein Relations Tell Us 1. The rate of spontaneous emission A21 is proportional to υ 3. At higher frequencies A21 >> B(υ) and all emission is spontaneous. A 21 = 8π hν3 c 3 B(ν) 2. Although

More information

( ) x10 8 m. The energy in a mole of 400 nm photons is calculated by: ' & sec( ) ( & % ) 6.022x10 23 photons' E = h! = hc & 6.

( ) x10 8 m. The energy in a mole of 400 nm photons is calculated by: ' & sec( ) ( & % ) 6.022x10 23 photons' E = h! = hc & 6. Introduction to Spectroscopy Spectroscopic techniques are widely used to detect molecules, to measure the concentration of a species in solution, and to determine molecular structure. For proteins, most

More information

RYDBERG STATES (6s, 6s ) OF METHYL AND

RYDBERG STATES (6s, 6s ) OF METHYL AND Laser Chem., Vol. 13, pp. 151-157 Reprints available directly from the Publisher Photocopying permitted by license only (C) 1993 Harwood Academic Publishers GmbH Printed in Malaysia 2+1 (2+2) REMPI-TOF

More information

Spectroscopy in frequency and time domains

Spectroscopy in frequency and time domains 5.35 Module 1 Lecture Summary Fall 1 Spectroscopy in frequency and time domains Last time we introduced spectroscopy and spectroscopic measurement. I. Emphasized that both quantum and classical views of

More information

Photoelectron spectroscopy via the 1 1 u state of diacetylene

Photoelectron spectroscopy via the 1 1 u state of diacetylene JOURNAL OF CHEMICAL PHYSICS VOLUME 116, NUMBER 10 8 MARCH 2002 Photoelectron spectroscopy via the 1 1 u state of diacetylene C. Ramos, P. R. Winter, and T. S. Zwier Department of Chemistry, Purdue University,

More information

VIBRATION-ROTATION SPECTRUM OF CO

VIBRATION-ROTATION SPECTRUM OF CO Rice University Physics 332 VIBRATION-ROTATION SPECTRUM OF CO I. INTRODUCTION...2 II. THEORETICAL CONSIDERATIONS...3 III. MEASUREMENTS...8 IV. ANALYSIS...9 April 2011 I. Introduction Optical spectroscopy

More information

Vibrational Autoionization in Polyatomic molecules

Vibrational Autoionization in Polyatomic molecules Vibrational Autoionization in Polyatomic molecules S.T. Pratt Annu. Rev. Phys. Chem. 2005. 56:281-308 2006. 12. 4. Choi, Sunyoung 1 Schedule 12/4 (Mon) - Introduction - Theoretical background 12/6 (Wed)

More information

CD Basis Set of Spectra that is used is that derived from comparing the spectra of globular proteins whose secondary structures are known from X-ray

CD Basis Set of Spectra that is used is that derived from comparing the spectra of globular proteins whose secondary structures are known from X-ray CD Basis Set of Spectra that is used is that derived from comparing the spectra of globular proteins whose secondary structures are known from X-ray crystallography An example of the use of CD Modeling

More information

Ultraviolet-Visible and Infrared Spectrophotometry

Ultraviolet-Visible and Infrared Spectrophotometry Ultraviolet-Visible and Infrared Spectrophotometry Ahmad Aqel Ifseisi Assistant Professor of Analytical Chemistry College of Science, Department of Chemistry King Saud University P.O. Box 2455 Riyadh 11451

More information

Linear and nonlinear spectroscopy

Linear and nonlinear spectroscopy Linear and nonlinear spectroscopy We ve seen that we can determine molecular frequencies and dephasing rates (for electronic, vibrational, or spin degrees of freedom) from frequency-domain or timedomain

More information

Quantum Chemistry. NC State University. Lecture 5. The electronic structure of molecules Absorption spectroscopy Fluorescence spectroscopy

Quantum Chemistry. NC State University. Lecture 5. The electronic structure of molecules Absorption spectroscopy Fluorescence spectroscopy Quantum Chemistry Lecture 5 The electronic structure of molecules Absorption spectroscopy Fluorescence spectroscopy NC State University 3.5 Selective absorption and emission by atmospheric gases (source:

More information

Lecture 3: Light absorbance

Lecture 3: Light absorbance Lecture 3: Light absorbance Perturbation Response 1 Light in Chemistry Light Response 0-3 Absorbance spectrum of benzene 2 Absorption Visible Light in Chemistry S 2 S 1 Fluorescence http://www.microscopyu.com

More information

Excited State Processes

Excited State Processes Excited State Processes Photophysics Fluorescence (singlet state emission) Phosphorescence (triplet state emission) Internal conversion (transition to singlet gr. state) Intersystem crossing (transition

More information

Experiment 6: Vibronic Absorption Spectrum of Molecular Iodine

Experiment 6: Vibronic Absorption Spectrum of Molecular Iodine Experiment 6: Vibronic Absorption Spectrum of Molecular Iodine We have already seen that molecules can rotate and bonds can vibrate with characteristic energies, each energy being associated with a particular

More information

Spectroscopic Investigation of Polycyclic Aromatic Hydrocarbons Trapped in Liquid Helium Clusters

Spectroscopic Investigation of Polycyclic Aromatic Hydrocarbons Trapped in Liquid Helium Clusters Spectroscopic Investigation of Polycyclic Aromatic Hydrocarbons Trapped in Liquid Helium Clusters Friedrich Huisken and Serge Krasnokutski Max-Planck-Institut für Strömungsforschung, Bunsenstr. 10, D-37073

More information

Introduction to Sources: Radiative Processes and Population Inversion in Atoms, Molecules, and Semiconductors Atoms and Molecules

Introduction to Sources: Radiative Processes and Population Inversion in Atoms, Molecules, and Semiconductors Atoms and Molecules OPTI 500 DEF, Spring 2012, Lecture 2 Introduction to Sources: Radiative Processes and Population Inversion in Atoms, Molecules, and Semiconductors Atoms and Molecules Energy Levels Every atom or molecule

More information

Today: general condition for threshold operation physics of atomic, vibrational, rotational gain media intro to the Lorentz model

Today: general condition for threshold operation physics of atomic, vibrational, rotational gain media intro to the Lorentz model Today: general condition for threshold operation physics of atomic, vibrational, rotational gain media intro to the Lorentz model Laser operation Simplified energy conversion processes in a laser medium:

More information

I 2 Vapor Absorption Experiment and Determination of Bond Dissociation Energy.

I 2 Vapor Absorption Experiment and Determination of Bond Dissociation Energy. I 2 Vapor Absorption Experiment and Determination of Bond Dissociation Energy. What determines the UV-Vis (i.e., electronic transitions) band appearance? Usually described by HOMO LUMO electron jump LUMO

More information

Luminescence. Photoluminescence (PL) is luminescence that results from optically exciting a sample.

Luminescence. Photoluminescence (PL) is luminescence that results from optically exciting a sample. Luminescence Topics Radiative transitions between electronic states Absorption and Light emission (spontaneous, stimulated) Excitons (singlets and triplets) Franck-Condon shift(stokes shift) and vibrational

More information

Fluorescence (Notes 16)

Fluorescence (Notes 16) Fluorescence - 2014 (Notes 16) XV 74 Jablonski diagram Where does the energy go? Can be viewed like multistep kinetic pathway 1) Excite system through A Absorbance S 0 S n Excite from ground excited singlet

More information

two slits and 5 slits

two slits and 5 slits Electronic Spectroscopy 2015January19 1 1. UV-vis spectrometer 1.1. Grating spectrometer 1.2. Single slit: 1.2.1. I diffracted intensity at relative to un-diffracted beam 1.2.2. I - intensity of light

More information

Molecular Luminescence. Absorption Instrumentation. UV absorption spectrum. lg ε. A = εbc. monochromator. light source. Rotating mirror (beam chopper)

Molecular Luminescence. Absorption Instrumentation. UV absorption spectrum. lg ε. A = εbc. monochromator. light source. Rotating mirror (beam chopper) Molecular Luminescence Absorption Instrumentation light source I 0 sample I detector light source Rotating mirror (beam chopper) motor b sample I detector reference I 0 UV absorption spectrum lg ε A =

More information

Chapter 4 Scintillation Detectors

Chapter 4 Scintillation Detectors Med Phys 4RA3, 4RB3/6R03 Radioisotopes and Radiation Methodology 4-1 4.1. Basic principle of the scintillator Chapter 4 Scintillation Detectors Scintillator Light sensor Ionizing radiation Light (visible,

More information

16. Reactions of the Radical Pairs. The I P Step of the Paradigm

16. Reactions of the Radical Pairs. The I P Step of the Paradigm 16. Reactions of the Radical Pairs. The I P Step of the Paradigm Let us consider the product forming step in the general photochemical paradigm, i.e., the I P process shown in Figure 1. From the discussion

More information

Chem Homework Set Answers

Chem Homework Set Answers Chem 310 th 4 Homework Set Answers 1. Cyclohexanone has a strong infrared absorption peak at a wavelength of 5.86 µm. (a) Convert the wavelength to wavenumber.!6!1 8* = 1/8 = (1/5.86 µm)(1 µm/10 m)(1 m/100

More information

Model Answer (Paper code: AR-7112) M. Sc. (Physics) IV Semester Paper I: Laser Physics and Spectroscopy

Model Answer (Paper code: AR-7112) M. Sc. (Physics) IV Semester Paper I: Laser Physics and Spectroscopy Model Answer (Paper code: AR-7112) M. Sc. (Physics) IV Semester Paper I: Laser Physics and Spectroscopy Section I Q1. Answer (i) (b) (ii) (d) (iii) (c) (iv) (c) (v) (a) (vi) (b) (vii) (b) (viii) (a) (ix)

More information

Assumed knowledge. Chemistry 2. Learning outcomes. Electronic spectroscopy of polyatomic molecules. Franck-Condon Principle (reprise)

Assumed knowledge. Chemistry 2. Learning outcomes. Electronic spectroscopy of polyatomic molecules. Franck-Condon Principle (reprise) Chemistry 2 Lecture 11 Electronic spectroscopy of polyatomic molecules Assumed knowledge For bound excited states, transitions to the individual vibrational levels of the excited state are observed with

More information

Born-Oppenheimer Approximation

Born-Oppenheimer Approximation Born-Oppenheimer Approximation Adiabatic Assumption: Nuclei move so much more slowly than electron that the electrons that the electrons are assumed to be obtained if the nuclear kinetic energy is ignored,

More information

Fundamentals of Spectroscopy for Optical Remote Sensing. Course Outline 2009

Fundamentals of Spectroscopy for Optical Remote Sensing. Course Outline 2009 Fundamentals of Spectroscopy for Optical Remote Sensing Course Outline 2009 Part I. Fundamentals of Quantum Mechanics Chapter 1. Concepts of Quantum and Experimental Facts 1.1. Blackbody Radiation and

More information

Excited States in Organic Light-Emitting Diodes

Excited States in Organic Light-Emitting Diodes Excited States in Organic Light-Emitting Diodes The metal-to-ligand charge transfer (MLCT) excited states of d 6 π coordination compounds have emerged as the most efficient for solar harvesting and sensitization

More information

Because light behaves like a wave, we can describe it in one of two ways by its wavelength or by its frequency.

Because light behaves like a wave, we can describe it in one of two ways by its wavelength or by its frequency. Light We can use different terms to describe light: Color Wavelength Frequency Light is composed of electromagnetic waves that travel through some medium. The properties of the medium determine how light

More information

Chapter 3. Electromagnetic Theory, Photons. and Light. Lecture 7

Chapter 3. Electromagnetic Theory, Photons. and Light. Lecture 7 Lecture 7 Chapter 3 Electromagnetic Theory, Photons. and Light Sources of light Emission of light by atoms The electromagnetic spectrum see supplementary material posted on the course website Electric

More information

I 2 Vapor Absorption Experiment and Determination of Bond Dissociation Energy.

I 2 Vapor Absorption Experiment and Determination of Bond Dissociation Energy. I 2 Vapor Absorption Experiment and Determination of Bond Dissociation Energy. What determines the UV-Vis (i.e., electronic transitions) band appearance? Usually described by HOMO LUMO electron jump LUMO

More information

Vibronic Coupling in Quantum Wires: Applications to Polydiacetylene

Vibronic Coupling in Quantum Wires: Applications to Polydiacetylene Vibronic Coupling in Quantum Wires: Applications to Polydiacetylene An Exhaustively Researched Report by Will Bassett and Cole Johnson Overall Goal In order to elucidate the absorbance spectra of different

More information

An Investigation of Benzene Using Ultrafast Laser Spectroscopy. Ryan Barnett. The Ohio State University

An Investigation of Benzene Using Ultrafast Laser Spectroscopy. Ryan Barnett. The Ohio State University An Investigation of Benzene Using Ultrafast Laser Spectroscopy Ryan Barnett The Ohio State University NSF/REU/OSU Advisor: Linn Van Woerkom Introduction Molecular spectroscopy has been used throughout

More information

EXPERIMENT 5. The Franck-Hertz Experiment (Critical Potentials) Introduction

EXPERIMENT 5. The Franck-Hertz Experiment (Critical Potentials) Introduction EXPERIMENT 5 The Franck-Hertz Experiment (Critical Potentials) Introduction In the early part of the twentieth century the structure of the atom was studied in depth. In the process of developing and refining

More information

The Franck-Hertz Experiment Physics 2150 Experiment No. 9 University of Colorado

The Franck-Hertz Experiment Physics 2150 Experiment No. 9 University of Colorado Experiment 9 1 Introduction The Franck-Hertz Experiment Physics 2150 Experiment No. 9 University of Colorado During the late nineteenth century, a great deal of evidence accumulated indicating that radiation

More information

wbt Λ = 0, 1, 2, 3, Eq. (7.63)

wbt Λ = 0, 1, 2, 3, Eq. (7.63) 7.2.2 Classification of Electronic States For all diatomic molecules the coupling approximation which best describes electronic states is analogous to the Russell- Saunders approximation in atoms The orbital

More information

Department of Chemistry Physical Chemistry Göteborg University

Department of Chemistry Physical Chemistry Göteborg University Department of Chemistry Physical Chemistry Göteborg University &RQVWUXFWLRQRIDSXOVHGG\HODVHU 3OHDVHREVHUYHWKDWWKHVDIHW\SUHFDXWLRQVRQSDJHPXVW EHIROORZHGRWKHUZLVHWKHUHLVDULVNRIH\HGDPDJH Renée Andersson -97,

More information

Chemistry Instrumental Analysis Lecture 3. Chem 4631

Chemistry Instrumental Analysis Lecture 3. Chem 4631 Chemistry 4631 Instrumental Analysis Lecture 3 Quantum Transitions The energy of a photon can also be transferred to an elementary particle by adsorption if the energy of the photon exactly matches the

More information

MS/MS .LQGVRI0606([SHULPHQWV

MS/MS .LQGVRI0606([SHULPHQWV 0DVV6SHFWURPHWHUV Tandem Mass Spectrometry (MS/MS) :KDWLV0606" Mass spectrometers are commonly combined with separation devices such as gas chromatographs (GC) and liquid chromatographs (LC). The GC or

More information

5.80 Small-Molecule Spectroscopy and Dynamics

5.80 Small-Molecule Spectroscopy and Dynamics MIT OpenCourseWare http://ocw.mit.edu 5.80 Small-Molecule Spectroscopy and Dynamics Fall 2008 For information about citing these materials or our Terms of Use, visit: http://ocw.mit.edu/terms. Lectures

More information

Lecture 0. NC State University

Lecture 0. NC State University Chemistry 736 Lecture 0 Overview NC State University Overview of Spectroscopy Electronic states and energies Transitions between states Absorption and emission Electronic spectroscopy Instrumentation Concepts

More information

V( x) = V( 0) + dv. V( x) = 1 2

V( x) = V( 0) + dv. V( x) = 1 2 Spectroscopy 1: rotational and vibrational spectra The vibrations of diatomic molecules Molecular vibrations Consider a typical potential energy curve for a diatomic molecule. In regions close to R e (at

More information

Laser Physics OXFORD UNIVERSITY PRESS SIMON HOOKER COLIN WEBB. and. Department of Physics, University of Oxford

Laser Physics OXFORD UNIVERSITY PRESS SIMON HOOKER COLIN WEBB. and. Department of Physics, University of Oxford Laser Physics SIMON HOOKER and COLIN WEBB Department of Physics, University of Oxford OXFORD UNIVERSITY PRESS Contents 1 Introduction 1.1 The laser 1.2 Electromagnetic radiation in a closed cavity 1.2.1

More information

CHEM Outline (Part 15) - Luminescence 2013

CHEM Outline (Part 15) - Luminescence 2013 CHEM 524 -- Outline (Part 15) - Luminescence 2013 XI. Molecular Luminescence Spectra (Chapter 15) Kinetic process, competing pathways fluorescence, phosphorescence, non-radiative decay Jablonski diagram

More information

Γ43 γ. Pump Γ31 Γ32 Γ42 Γ41

Γ43 γ. Pump Γ31 Γ32 Γ42 Γ41 Supplementary Figure γ 4 Δ+δe Γ34 Γ43 γ 3 Δ Ω3,4 Pump Ω3,4, Ω3 Γ3 Γ3 Γ4 Γ4 Γ Γ Supplementary Figure Schematic picture of theoretical model: The picture shows a schematic representation of the theoretical

More information

Spectroscopy Problem Set February 22, 2018

Spectroscopy Problem Set February 22, 2018 Spectroscopy Problem Set February, 018 4 3 5 1 6 7 8 1. In the diagram above which of the following represent vibrational relaxations? 1. Which of the following represent an absorbance? 3. Which of following

More information

5.74 Introductory Quantum Mechanics II

5.74 Introductory Quantum Mechanics II MIT OpenCourseWare http://ocw.mit.edu 5.74 Introductory Quantum Mechanics II Spring 2009 For information about citing these materials or our Terms of Use, visit: http://ocw.mit.edu/terms. p. 10-0 10..

More information

Richard Miles and Arthur Dogariu. Mechanical and Aerospace Engineering Princeton University, Princeton, NJ 08540, USA

Richard Miles and Arthur Dogariu. Mechanical and Aerospace Engineering Princeton University, Princeton, NJ 08540, USA Richard Miles and Arthur Dogariu Mechanical and Aerospace Engineering Princeton University, Princeton, NJ 08540, USA Workshop on Oxygen Plasma Kinetics Sept 20, 2016 Financial support: ONR and MetroLaser

More information

CHAPTER 7 SUMMARY OF THE PRESENT WORK AND SUGGESTIONS FOR FUTURE WORK

CHAPTER 7 SUMMARY OF THE PRESENT WORK AND SUGGESTIONS FOR FUTURE WORK 161 CHAPTER 7 SUMMARY OF THE PRESENT WORK AND SUGGESTIONS FOR FUTURE WORK 7.1 SUMMARY OF THE PRESENT WORK Nonlinear optical materials are required in a wide range of important applications, such as optical

More information

Confocal Microscopy Imaging of Single Emitter Fluorescence and Hanbury Brown and Twiss Photon Antibunching Setup

Confocal Microscopy Imaging of Single Emitter Fluorescence and Hanbury Brown and Twiss Photon Antibunching Setup 1 Confocal Microscopy Imaging of Single Emitter Fluorescence and Hanbury Brown and Twiss Photon Antibunching Setup Abstract Jacob Begis The purpose of this lab was to prove that a source of light can be

More information

Quantum Master Equations for the Electron Transfer Problem

Quantum Master Equations for the Electron Transfer Problem 20/01/2010 Quantum Master Equations for the Electron Transfer Problem Seminarvortrag Dekohaerenz und Dissipation in Quantensystemen Antonio A. Gentile The general transport problem in micro/mesoscopic

More information

Reflection = EM strikes a boundary between two media differing in η and bounces back

Reflection = EM strikes a boundary between two media differing in η and bounces back Reflection = EM strikes a boundary between two media differing in η and bounces back Incident ray θ 1 θ 2 Reflected ray Medium 1 (air) η = 1.00 Medium 2 (glass) η = 1.50 Specular reflection = situation

More information

Skoog Chapter 6 Introduction to Spectrometric Methods

Skoog Chapter 6 Introduction to Spectrometric Methods Skoog Chapter 6 Introduction to Spectrometric Methods General Properties of Electromagnetic Radiation (EM) Wave Properties of EM Quantum Mechanical Properties of EM Quantitative Aspects of Spectrochemical

More information

Rotational Raman Spectroscopy

Rotational Raman Spectroscopy Rotational Raman Spectroscopy If EM radiation falls upon an atom or molecule, it may be absorbed if the energy of the radiation corresponds to the separation of two energy levels of the atoms or molecules.

More information

Femtosecond photoelectron spectroscopy of I 2 Ar n clusters n 6,9,12,16,20

Femtosecond photoelectron spectroscopy of I 2 Ar n clusters n 6,9,12,16,20 JOURNAL OF CHEMICAL PHYSICS VOLUME 111, NUMBER 23 15 DECEMBER 1999 Femtosecond photoelectron spectroscopy of I 2 Ar n clusters n 6,9,12,16,20 B. Jefferys Greenblatt, a) Martin T. Zanni, and Daniel M. Neumark

More information

Chapter-4 Stimulated emission devices LASERS

Chapter-4 Stimulated emission devices LASERS Semiconductor Laser Diodes Chapter-4 Stimulated emission devices LASERS The Road Ahead Lasers Basic Principles Applications Gas Lasers Semiconductor Lasers Semiconductor Lasers in Optical Networks Improvement

More information

Single Photon detectors

Single Photon detectors Single Photon detectors Outline Motivation for single photon detection Semiconductor; general knowledge and important background Photon detectors: internal and external photoeffect Properties of semiconductor

More information

Chapter 28. Atomic Physics

Chapter 28. Atomic Physics Chapter 28 Atomic Physics Quantum Numbers and Atomic Structure The characteristic wavelengths emitted by a hot gas can be understood using quantum numbers. No two electrons can have the same set of quantum

More information

CHAPTER 3 RESULTS AND DISCUSSION

CHAPTER 3 RESULTS AND DISCUSSION CHAPTER 3 RESULTS AND DISCUSSION 3.1 CHAPTER OUTLINE This chapter presents the data obtained from the investigation of each of the following possible explanations: (1) Experimental artifacts. (2) Direct

More information

Last Updated: April 22, 2012 at 7:49pm

Last Updated: April 22, 2012 at 7:49pm Page 1 Electronic Properties of d 6 π Coordination Compounds The metal-to-ligand charge transfer (MLCT) excited states of d 6 π coordination compounds have emerged as the most efficient for both solar

More information

Wolfgang Demtroder. Molecular Physics. Theoretical Principles and Experimental Methods WILEY- VCH. WILEY-VCH Verlag GmbH & Co.

Wolfgang Demtroder. Molecular Physics. Theoretical Principles and Experimental Methods WILEY- VCH. WILEY-VCH Verlag GmbH & Co. Wolfgang Demtroder Molecular Physics Theoretical Principles and Experimental Methods WILEY- VCH WILEY-VCH Verlag GmbH & Co. KGaA v Preface xiii 1 Introduction 1 1.1 Short Historical Overview 2 1.2 Molecular

More information

Molecular spectroscopy

Molecular spectroscopy Molecular spectroscopy Origin of spectral lines = absorption, emission and scattering of a photon when the energy of a molecule changes: rad( ) M M * rad( ' ) ' v' 0 0 absorption( ) emission ( ) scattering

More information

Atomic Spectra. d sin θ = mλ (1)

Atomic Spectra. d sin θ = mλ (1) Atomic Spectra Objectives: To measure the wavelengths of visible light emitted by atomic hydrogen and verify that the measured wavelengths obey the empirical Rydberg formula. To observe emission spectra

More information

Interaction of Molecules with Radiation

Interaction of Molecules with Radiation 3 Interaction of Molecules with Radiation Atoms and molecules can exist in many states that are different with respect to the electron configuration, angular momentum, parity, and energy. Transitions between

More information

CHAPTER A2 LASER DESORPTION IONIZATION AND MALDI

CHAPTER A2 LASER DESORPTION IONIZATION AND MALDI Back to Basics Section A: Ionization Processes CHAPTER A2 LASER DESORPTION IONIZATION AND MALDI TABLE OF CONTENTS Quick Guide...27 Summary...29 The Ionization Process...31 Other Considerations on Laser

More information

MOLECULAR SPECTROSCOPY

MOLECULAR SPECTROSCOPY MOLECULAR SPECTROSCOPY First Edition Jeanne L. McHale University of Idaho PRENTICE HALL, Upper Saddle River, New Jersey 07458 CONTENTS PREFACE xiii 1 INTRODUCTION AND REVIEW 1 1.1 Historical Perspective

More information

Chemistry 126 Spectroscopy Week # 5 Diatomic Bonding/Electronic Structure & the Franck-Condon Approximation As well described in Chapter 11 of

Chemistry 126 Spectroscopy Week # 5 Diatomic Bonding/Electronic Structure & the Franck-Condon Approximation As well described in Chapter 11 of Chemistry 126 Spectroscopy Week # 5 Diatomic Bonding/Electronic Structure & the Franck-Condon Approximation As well described in Chapter 11 of McQuarrie, for diatomics heavier than 2, we start with the

More information

MODERN OPTICS. P47 Optics: Unit 9

MODERN OPTICS. P47 Optics: Unit 9 MODERN OPTICS P47 Optics: Unit 9 Course Outline Unit 1: Electromagnetic Waves Unit 2: Interaction with Matter Unit 3: Geometric Optics Unit 4: Superposition of Waves Unit 5: Polarization Unit 6: Interference

More information

P. W. Atkins and R. S. Friedman. Molecular Quantum Mechanics THIRD EDITION

P. W. Atkins and R. S. Friedman. Molecular Quantum Mechanics THIRD EDITION P. W. Atkins and R. S. Friedman Molecular Quantum Mechanics THIRD EDITION Oxford New York Tokyo OXFORD UNIVERSITY PRESS 1997 Introduction and orientation 1 Black-body radiation 1 Heat capacities 2 The

More information

Quantum control of dissipative systems. 1 Density operators and mixed quantum states

Quantum control of dissipative systems. 1 Density operators and mixed quantum states Quantum control of dissipative systems S. G. Schirmer and A. I. Solomon Quantum Processes Group, The Open University Milton Keynes, MK7 6AA, United Kingdom S.G.Schirmer@open.ac.uk, A.I.Solomon@open.ac.uk

More information

Mass Analyzers. Principles of the three most common types magnetic sector, quadrupole and time of flight - will be discussed herein.

Mass Analyzers. Principles of the three most common types magnetic sector, quadrupole and time of flight - will be discussed herein. Mass Analyzers After the production of ions in ion sources, the next critical step in mass spectrometry is to separate these gas phase ions according to their mass-to-charge ratio (m/z). Ions are extracted

More information

Pulsed Lasers Revised: 2/12/14 15: , Henry Zmuda Set 5a Pulsed Lasers

Pulsed Lasers Revised: 2/12/14 15: , Henry Zmuda Set 5a Pulsed Lasers Pulsed Lasers Revised: 2/12/14 15:27 2014, Henry Zmuda Set 5a Pulsed Lasers 1 Laser Dynamics Puled Lasers More efficient pulsing schemes are based on turning the laser itself on and off by means of an

More information

B 2 P 2, which implies that g B should be

B 2 P 2, which implies that g B should be Enhanced Summary of G.P. Agrawal Nonlinear Fiber Optics (3rd ed) Chapter 9 on SBS Stimulated Brillouin scattering is a nonlinear three-wave interaction between a forward-going laser pump beam P, a forward-going

More information

Chap. 12 Photochemistry

Chap. 12 Photochemistry Chap. 12 Photochemistry Photochemical processes Jablonski diagram 2nd singlet excited state 3rd triplet excited state 1st singlet excited state 2nd triplet excited state 1st triplet excited state Ground

More information

A study of the predissociation of NaK molecules in the 6 1 state by optical optical double resonance spectroscopy

A study of the predissociation of NaK molecules in the 6 1 state by optical optical double resonance spectroscopy A study of the predissociation of NaK molecules in the 6 1 state by optical optical double resonance spectroscopy Z. J. Jabbour a) and J. Huennekens Department of Physics, Lehigh University, Bethlehem,

More information

Introduction to Fluorescence Spectroscopies I. Theory

Introduction to Fluorescence Spectroscopies I. Theory March 22, 2006; Revised January 26, 2011 for CHMY 374 Adapted for CHMY 362 November 13, 2011 Edited by Lauren Woods December 2, 2011 17mar14, P.Callis; 1feb17 P. Callis, 29jan18 P. Callis Introduction

More information

Lecture- 08 Emission and absorption spectra

Lecture- 08 Emission and absorption spectra Atomic and Molecular Absorption Spectrometry for Pollution Monitoring Dr. J R Mudakavi Department of Chemical Engineering Indian Institute of Science, Bangalore Lecture- 08 Emission and absorption spectra

More information

Chapter 17: Fundamentals of Spectrophotometry

Chapter 17: Fundamentals of Spectrophotometry Chapter 17: Fundamentals of Spectrophotometry Spectroscopy: the science that deals with interactions of matter with electromagnetic radiation or other forms energy acoustic waves, beams of particles such

More information

MIT Department of Nuclear Science & Engineering

MIT Department of Nuclear Science & Engineering 1 MIT Department of Nuclear Science & Engineering Thesis Prospectus for the Bachelor of Science Degree in Nuclear Science and Engineering Nicolas Lopez Development of a Nanoscale Magnetometer Through Utilization

More information

Chemistry 524--Final Exam--Keiderling May 4, :30 -?? pm SES

Chemistry 524--Final Exam--Keiderling May 4, :30 -?? pm SES Chemistry 524--Final Exam--Keiderling May 4, 2011 3:30 -?? pm -- 4286 SES Please answer all questions in the answer book provided. Calculators, rulers, pens and pencils are permitted. No open books or

More information

Saturation Absorption Spectroscopy of Rubidium Atom

Saturation Absorption Spectroscopy of Rubidium Atom Saturation Absorption Spectroscopy of Rubidium Atom Jayash Panigrahi August 17, 2013 Abstract Saturated absorption spectroscopy has various application in laser cooling which have many relevant uses in

More information